High-Performance Electrochromic Devices Based ... - ACS Publications

May 13, 2016 - CIQ, Departamento de Química e. Bioquímica, Faculdade de Ciências, Universidade do Porto, 4169-007 Porto, Portugal. §. CeNTI, Rua ...
1 downloads 0 Views 5MB Size
Research Article www.acsami.org

High-Performance Electrochromic Devices Based on Poly[Ni(salen)]Type Polymer Films Marta Nunes,† Mariana Araújo,† Joana Fonseca,§ Cosme Moura,‡ Robert Hillman,∥ and Cristina Freire*,† †

REQUIMTE/LAQV, Departamento de Química e Bioquímica, Faculdade de Ciências, and ‡CIQ, Departamento de Química e Bioquímica, Faculdade de Ciências, Universidade do Porto, 4169-007 Porto, Portugal § CeNTI, Rua Fernando Mesquita, 2785, 4760-034 Vila Nova de Famalicão, Portugal ∥ Department of Chemistry, University of Leicester, Leicester LE1 7RH, United Kingdom S Supporting Information *

ABSTRACT: We report the application of two poly[Ni(salen)]-type electroactive polymer films as new electrochromic materials. The two films, poly[Ni(3-Mesalen)] (poly[1]) and poly[Ni(3-MesaltMe)] (poly[2]), were successfully electrodeposited onto ITO/PET flexible substrates, and their voltammetric characterization revealed that poly[1] showed similar redox profiles in LiClO4/CH3CN and LiClO4/propylene carbonate (PC), while poly[2] showed solvent-dependent electrochemical responses. Both films showed multielectrochromic behavior, exhibiting yellow, green, and russet colors according to their oxidation state, and promising electrochromic properties with high electrochemical stability in LiClO4/PC supporting electrolyte. In particular, poly[1] exhibited a very good electrochemical stability, changing color between yellow and green (λ = 750 nm) during 9000 redox cycles, with a charge loss of 34.3%, an optical contrast of ΔT = 26.2%, and an optical density of ΔOD = 0.49, with a coloration efficiency of η = 75.55 cm2 C−1. On the other hand, poly[2] showed good optical contrast for the color change from green to russet (ΔT = 58.5%), although with moderate electrochemical stability. Finally, poly[1] was used to fabricate a solid-state electrochromic device using lateral configuration with two figures of merit: a simple shape (typology 1) and a butterfly shape (typology 2); typology 1 showed the best performance with optical contrast ΔT = 88.7% (at λ = 750 nm), coloration efficiency η = 130.4 cm2 C−1, and charge loss of 37.0% upon 3000 redox cycles. KEYWORDS: conducting polymers, metal salen complexes, spectroelectrochemistry, electrochromism, solid-state electrochromic cell

1. INTRODUCTION

potential, the color change occurs as a result of the charge/ discharge of the cell.14,15 Among the several known EC materials (e.g., Prussian Blue,16 WO3,17 and viologens18), conducting polymers (CPs)19 have attracted particular attention due to their outstanding coloration efficiency, fast switching times, multichromism, high optical and electrochemical stability, thin film mechanical flexibility, and cost effectiveness.20−22 In fact, several works have reported the EC behavior of the polypyrrole,23−25 polyaniline (PANI),26 polythiophene,27,28 and their derivatives, as well as the fabrication of efficient complementary ECDs comprising CPs (e.g., polythiophene-derivatives@PANI29,30 and polypyrrole@polythiophene-derivatives31). Typically, the electric conductivity of these CPs is very high, but can be limited by the existence of structural defects, which may interrupt the mobility of charge carriers. A new class of conducting polymers can be designed by incorporation of transition metals as functional groups or as units within the

The focus on electrochromic (EC) materials started after the pioneering work done by Deb1 in 1969 on electrochromism, which occurs when electroactive species exhibit reversible change or bleaching of color, upon their electrochemical oxidation/reduction.2−4 Currently, EC materials are mainly applied in smart windows for green architecture5 and in antiglare rear view mirrors.6 Other proposed applications are in displays of portable and flexible electronic devices, including smart cards, reusable price labels and electronic papers,7−9 protective eyewear,10 and in textiles for adaptative camouflage or simply for fashion.11 To take advantage of the special properties of the EC materials in these practical and commercial applications, it is necessary to achieve their suitable integration into solid-state electrochromic cells (electrochromic devices, ECDs). The prototype ECDs are generally fabricated using optically transparent electrodes, in a “sandwich”12 or lateral13 configuration, and consist of two electrodes (in which at least one is composed of an EC material), separated or covered by a layer of electrolyte.6 With the application of an appropriate electrical © 2016 American Chemical Society

Received: February 16, 2016 Accepted: May 13, 2016 Published: May 13, 2016 14231

DOI: 10.1021/acsami.6b01977 ACS Appl. Mater. Interfaces 2016, 8, 14231−14243

Research Article

ACS Applied Materials & Interfaces

(II), [Ni(3-MesaltMe)] [2], and respective salen ligands were prepared by methods described in the literature.39 Acetonitrile and propylene carbonate (PC) (Romil, pro analysis grade), LiClO4 (Aldrich, purity 99%), and poly(methyl methacrylate) (PMMA, Aldrich) were used as received. The X-ray photoelectron spectroscopy (XPS) measurements were performed at CEMUP (Porto, Portugal), in a VG Scientific ESCALAB 200A spectrometer using nonmonochromatized Al Kα radiation (1486.6 eV). The XPS spectra were deconvoluted with the XPSPEAK 4.1 software, using a nonlinear least-squares fitting routine after a Shirley-type background subtraction. To correct possible deviations caused by sample charging, the C 1s band at 284.6 eV was taken as internal standard. The surface atom percentages were calculated from the corresponding peak areas, using sensitivity factors provided by the manufacturer. Scanning electron microscopy/energy-dispersive X-ray spectroscopy (SEM/EDS) analyses were performed at CEMUP (Porto, Portugal) in a high-resolution environmental scanning electron microscope FEI Quanta 400 FEG ESEM, attached to an EDAX Genesis X4M X-ray spectrometer. The micrographs were obtained in secondary and backscattered electron modes. Electrochemical studies were performed using an Autolab PGSTAT 30 potentiostat/galvanostat (EcoChimie B.V.), controlled by a GPES software. Studies in solution were performed on a three-electrode and single or separate compartment cells, enclosed in a grounded Faraday cage, using an Ag/AgCl (NaCl/1.0 mol dm−3) electrode (Metrohm ref 6.0724.140) as the reference electrode, a Pt wire or a Pt grid (in the case of separate compartment cell) as the counter electrode, and poly(ethylene terephthalate) (PET) coated with indium tin oxide (ITO) (ITO/PET) (Aldrich, resistivity of 60 Ω sq−1) as the working electrode. Coulometric studies used a Pt disk working electrode (area 0.0314 cm2, BAS), which was previously polished with aluminum oxide of particle size 0.3 μm (Buehler) on a microcloth polishing pad (Buehler), then washed with ultrapure water (resistivity 18.2 MΩ cm at 25 °C, Millipore) and CH3CN. The spectroelectrochemical studies were performed in situ using an Agilent 8453 spectrophotometer (with diode array detection) coupled to the potentiostat/galvanostat. The experimental apparatus comprised a Teflon cell with an Ag/AgCl (NaCl/3.0 mol dm−3) (Bio-Logic) reference electrode, a Pt grid counter electrode, and ITO/PET (typical area 0.785 cm2) as working electrode. Particularly, in the ECD monitoring was used a PerkinElmer Lambda 35 UV/vis spectrometer coupled to an Autolab PGSTAT 302N potentiostat/galvanostat (EcoChimie B.V.). 2.2. Preparation and Characterization of Polymeric Films. The poly[Ni(salen)] films were deposited potentiodynamically from corresponding monomer solutions (1.0 mmol dm−3 [Ni(salen)] in 0.1 mol dm−3 LiClO4/CH3CN), by cycling the potential of the working electrode (ITO/PET 3.0 cm2) between 0.0 and 1.3 V, at the scan rate of 0.02 V s−1, during 10 cycles; other conditions are specified in the corresponding text and figure legends. After film deposition, the modified electrodes were rinsed with CH3CN, immersed in monomer-free 0.1 mol dm−3 LiClO4/CH3CN or LiClO4/PC solutions, and cycled voltammetrically in the potential ranges 0.0−1.3 or 0.0−1.4 V, at 0.020 V s−1. The electroactive surface coverage, Γ/μmol cm−2, of each film was determined by a coulometric assay,32,40 using cyclic voltammograms obtained in monomer-free 0.1 mol dm−3 LiClO4/CH3CN solution at a scan rate v = 0.01 V s−1, to ensure complete film redox conversion. The doping level (n) values used for Γ determination were calculated from comparison of coulometric data for films deposition and cycling in 0.1 mol dm−3 LiClO4/CH3CN solution with the Pt electrode, as described in the literature,32,40 considering that the obtained n-values are valid for other electrode substrates; the n-values were found to be n = 0.69 and 0.08 for poly[1] and poly[2], respectively. Although the values are very different from each other, this is a recognized behavior for these systems, and both values are in the range known for poly[M(salen)] films;32,34 their difference cannot be rationalized in terms of monomer ligand structures, which are very similar. Consequently, the supra-

polymer backbone. The latter can be achieved by the preparation of poly[M(salen)]-type films (M = transition metal).32 The electroactive poly[M(salen)] metallo-polymers films are formed through the oxidative polymerization of transition metal complexes with salen-type ligands onto the electrode surface,33 and exhibit poly(phenylene)-like properties, with the transition cations acting as a bridge between the phenyl rings.34 Their potential application as EC materials has been explored by Freire’s group.35,36 The colorimetric properties of polymers based on salen-type complexes of Cu(II), Ni(II), and Pd(II) deposited over transparent flexible electrodes of polyethylene terephthalate coated with indium tin oxide (ITO/PET) were first studied by Pinheiro et al.,35 using the CIELAB coordinates and spectroelectrochemical studies. More recently, Branco et al.36 studied other series of [M(salen)]-derived electroactive films, and selected materials were chosen to build homemade ECDs, whose preliminary performance was also assessed. The reported results addressed some promising EC properties for the poly[M(salen)] films in terms of color changes and optical contrasts, but they showed moderate electrochemical stabilities, which is one of the drawbacks for technological applications. Nevertheless, the results showed that the synthetic versatility of the metal salen complexes (by the introduction of different substituents into the salen skeleton or by using different metal centers) allowed fine-tuning of the final EC features of the resulting polymeric systems, providing important guidelines for film molecular design optimization. These observations motivated us to more detailed studies to achieve other poly[M(salen)]-type films with improved EC performance and electrochemical stability. Consequently, in this Article, we report the preparation and study of two novel electrochromic systems based on the electroactive polymers: poly[Ni(3-Mesalen)], N,N′-bis(3-methylsalicylideneiminate) nickel(II) (poly[1]), and poly[Ni(3-MesaltMe)], N,N′-2,3dimethylbutane-2,3-dyil-bis(3-methylsalicylideneiminate) nickel(II) (poly[2]); both salen ligands have a methyl group in the 3-position of each salicylaldehyde moiety and differ on the absence/presence of methyl substituents in the imine bridge, respectively. For the first time, the influence of two electrolyte solutions, LiClO4/CH3CN or LiClO4/PC, on the redox and optical properties of the films was evaluated. The studies allowed the full evaluation of EC properties (switching times, optical contrasts and densities, coloration efficiencies, and electrochemical stabilities) and were used as guidance to choose poly[1] as the polymeric film with the best EC properties for ECD fabrication with flexible substrates, using a lateral configuration with two figures of merit: simple lateral assemblage and butterfly shape. The former EC device showed the best promising EC performance and electrochemical stability: optical contrast ΔT = 88.7% (λ = 750 nm), coloration efficiency η = 130.4 cm2 C−1, and charge loss of 37.0% upon 3000 redox cycles. This work corresponds to a step forward in the design of EC materials and device fabrication based in poly[M(salen)] films, because poly[1] surpassed all of the previous EC metallopolymers analogues35,36 and are competitive with other EC materials20,37,38 for electronic devices, smart cards, or labels.

2. EXPERIMENTAL SECTION 2.1. Materials and Instrumentation. The complexes N,N′-bis(3methylsalicylideneiminate) nickel(II), [Ni(3-Mesalen)] [1], and N,N′2,3-dimethylbutane-2,3-dyil-bis(3-methylsalicylideneiminate) nickel14232

DOI: 10.1021/acsami.6b01977 ACS Appl. Mater. Interfaces 2016, 8, 14231−14243

Research Article

ACS Applied Materials & Interfaces molecular film structure may have an important key role in the degree of film doping, which is usually difficult to anticipate and quantify. The monitoring of the redox process by in situ UV−vis spectroscopy was performed with films prepared with 5 electrodeposition cycles. The films were subsequently cycled five times between −0.1 and 1.3 V, at v = 0.020 V s−1, using monomer-free LiClO4/CH3CN or LiClO4/PC 0.1 mol dm−3 solutions. The UV−vis spectra were acquired simultaneously at 0.5 s intervals, in the wavelength range of 315−1100 nm. The molar extinction coefficients, ε/cm−1 mol−1 dm3, of all observed electronic bands were estimated using a combination of the Beer−Lambert and Faraday laws (eq 1):32,40,41

Abs(λ) = ε(λ)Q /nFA

slow stirring, in the two-compartment cell. The electrolyte was a semisolid polymeric electrolyte based on PMMA, prepared by a procedure adapted from the literature.42 Briefly, a mixture of PMMA (7%), LiClO4 (3%), CH3CN (58%), and PC (32%) (% m/m) was stirred during about 20 h, at room temperature, to yield a translucent and malleable gel. We note that the high viscosity of this type of electrolyte makes it impractical for conventional electrochemical cell configurations (see above), but its use was explored here in device configuration as a consequence of its relevance to practical applications. The electrochromic devide (ECD) of typology 1 was characterized by chronoamperometry, during 3000 redox cycles, applying potential pulses of 200 s, with potential alternating between −1.0 and 1.0 V (color transition yellow ↔ green). During the first 1000 redox cycles, the experiment was monitored by UV−vis spectroscopy (chronoabsorptometry), at λ = 750 nm. The ECD of typology 2 was also characterized by chronoamperometry, but applying potential pulses of 100 s, with potential alternating between −1.1 and −0.25 V (yellow ↔ green). Note that these potential values are, in reality, potential differences between the two electrodes; because there is no reference electrode, potential values are not directly placed on an absolute scale.

(1)

where Q is the charge (C), n is the doping level, F is the Faraday constant, and A is the electrode area (cm2). Films for SEM/EDS and XPS analyses were prepared with 10 and 15 cycles, and the reduced and oxidized states were achieved by the application of 0.0 or 1.3 V, during 300 s, in a monomer-free LiClO4/ CH3CN 0.1 mol dm−3 solution, after the electrodeposition. For poly[1], we also studied by XPS the reduced and oxidized films in LiClO4/PC 0.1 mol dm−3. 2.3. Electrochromic Properties Evaluation. The EC properties of the films were explored by a double potential step method (chronoamperometry) coupled with UV−vis spectroscopy (chronoabsorptometry). The studies were performed in 0.1 mol dm−3 LiClO4/ CH3CN or LiClO4/PC solutions, considering two possible color transitions: yellow ↔ green and green ↔ russet. In the double potential step experiment, the potential was set at the initial potential value (0.0 or 0.7 V, according to the color change) for a period of time of 50 s and was stepped to a second potential (0.7, 1.3, or 1.15 V) for another period of 50 s, after being switched back to the initial potential. The simultaneous chronoabsorptometric measurements were performed at the fixed wavelengths of λ = 510 nm (green ↔ russet) and λ = 750 nm (yellow ↔ green), acquiring the UV−vis spectra at intervals of 1 s, during 4 redox cycles. The films used in these studies were electrodeposited using 5 potentiodynamic cycles. Electrochemical stability tests were performed by chronoamperometry, using the same experimental conditions, but over extended time intervals (400−9000 redox cycles, ca.11 h−11 days). The films used in these studies were prepared with 10 electrodeposition cycles. The change of the optical density, ΔOD (or optical absorbance change, ΔAbs), was estimated using eq 2:15

ΔAbs(λ) = ΔOD(λ) = log(Tred(λ)/Tox(λ))

3. RESULTS AND DISCUSSION 3.1. Electrochemical Preparation and Characterization of Polymeric Films. The voltammetric responses obtained during the electropolymerization of both [Ni(salen)] complexes are depicted in Figure 1, and the peak potential values are summarized in Table S1. In the first anodic half-cycle, two peaks are observed at Epa = 0.87 and 1.08 V and one peak at Epa = 0.99 V for [Ni(3Mesalen)] [1] and [Ni(3-MesaltMe)] [2] complexes, respectively, which are attributed to the oxidation of monomers, resulting in the formation of oligomers/polymer in the vicinity of the working electrode surface.32 In the reverse scan, two peaks are detected at Epc = 0.69 and 0.20 V for [Ni(3Mesalen)] and one peak at Epc = 0.64 with a slight inflection at Epc = 0.50 V for [Ni(3-MesaltMe)], which correspond to the reduction of film deposited in the preceding anodic scan.43,44 In the second and subsequent voltammetric cycles (5th cycle in Table S1), a new peak appears at Epa ≈ 0.42 V for the [Ni(3Mesalen)] system, which is assigned to polymer film oxidation deposited during the previous cycle(s);43,44 in the reverse cathodic, the peaks at Epc = 0.64 and 0.22 V are due to the oligomers and polymer film reductions. In the case of [Ni(3MesaltMe)], two new peaks appear during the second and subsequent scans (fifth cycle in Table S1) at Epa = 0.59 and 0.68 V, due to polymer film oxidation, and the peak attributed to the monomer oxidation shows now a slight inflection at Epa = 1.17 V; the second and the following cathodic scans are significantly different from the first one, showing three cathodic waves at Epc = 0.87 V (oligomers reduction) and Epc = 0.58 and 0.45 V, due to polymer film reduction. The growth of both polymers is proved by the increasing current intensity of the peaks corresponding to the oxidation and reduction of film in consecutive cycles, and by the visual inspection of a green film deposited on the electrode surface at the end of the process. The voltammetric responses of the as-prepared poly[1] and poly[2] films in the electrolyte solutions LiClO4/CH3CN and LiClO4/PC 0.1 mol dm−3 are depicted in Figure 2, and in Table S2 are summarized the potential peak values. It is possible to identify two anodic peaks at Epa = 0.39−0.88 V and Epa = 0.94−1.29 V and two cathodic peaks at Epc = 0.17− 0.41 V and Epc = 0.56−0.83 V. The results revealed that poly[1] film showed similar redox profiles either in LiClO4/CH3CN or

(2)

where Tred and Tox are the transmittance values of the films in reduced and oxidized states, at a fixed wavelength (λ = 510 or 750 nm). The coloration efficiency, η, is defined as the change of the optical density, ΔOD, and it was estimated by considering the total charge passed during the potential pulse, per unit area, Qd:12,15 η = ΔOD/Q d

(3)

2.4. Fabrication and Characterization of Electrochromic Devices. ECDs were built using as EC material poly[1] films electrodeposited on flexible substrates of ITO/PET (4.5 × 3.0 cm2). The devices were assembled in lateral configuration, with the electrodes side-by-side, in two different typologies: typology 1, with a simple shape, and typology 2, with a butterfly shape. Both electrodes are based on the same EC material, and, in case of the device of typology 2, one of the electrodes corresponds to the butterfly inside area, while the other corresponds to the butterfly outside area. Following film deposition, the ECDs were assembled by making a cut under the film and ITO surface, to define the limits of each electrode and prevent electrical contact between each part. The cut is a simple vertical line in case of ECD of typology 1 and corresponds to the butterfly shape for the device of typology 2. A layer of a PMMAbased electrolyte then was deposited over both electrodes. The poly[1] films were electrodeposited potentiodynamically in the potential range 0.3−1.5 V, at v = 0.020 V s−1, for 10 cycles and with 14233

DOI: 10.1021/acsami.6b01977 ACS Appl. Mater. Interfaces 2016, 8, 14231−14243

Research Article

ACS Applied Materials & Interfaces

by XPS. Surface atomic percentages and atomic ratios are summarized in Table 1 for poly[1] in LiClO4/CH3CN and LiClO4/PC and in Table S3 for poly[2] in LiClO4/CH3CN. For poly[1] and poly[2] reduced films in LiClO4/CH3CN, the XPS-derived atomic ratios N/Ni and O/Ni are slightly higher than those expected, based on the Ni:salen stoichiometric ratio, N/Ni = O/N = 2. Combined with the appearance of Cl, this indicates the presence of ClO4− and/or CH3CN at low levels, as a result of the supporting electrolyte trapping in polymeric matrixes during the deposition or cycling processes.32 In comparison, the oxidized poly[1] and poly[2] films showed higher N/Ni, O/Ni, and Cl/Ni atomic ratios. This significant increase (higher for O/Ni and Cl/Ni ratios) is explained by the presence of ClO4−/CH3CN inside the films, due to the charge compensation/solvation during the films oxidation, as observed previously for similar systems32 (deconvoluted XPS spectra in O 1s region for reduced and oxidized poly[1] in Figure S1). This assumption is supported by the increase of the Cl/O ratio in a proportion of ca. 1:4 (in agreement with ClO4− composition) from the reduced to the oxidized films. In LiClO4/PC, poly[1] film showed similar changes in atomic ratios on going from the reduced to the oxidized state, which can be interpreted as in LiClO4/CH3CN electrolyte solution. Furthermore, it is worth mentioning that there is a larger increase in O/Ni and Cl/Ni ratios, from the reduced to the oxidized state, in LiClO4/CH3CN (increase of 79.7%) than in LiClO4/PC (increase of 36.4%), suggesting a higher degree of ClO4− ingress during film oxidation in the former electrolyte solution. Another interesting aspect, when comparing poly[1] and poly[2] films in LiClO4/CH3CN electrolyte solution, is that both O/Ni and Cl/Ni ratios are lower in the reduced state and higher in the oxidized state of poly[1]. This indicates a lower % of Li+ClO4− entrapped within poly[1] film matrix relative to poly[2] in the reduced state, but a higher % of ClO4− occluded in the oxidized state; this may be to some extent a consequence of the higher n-doping value of poly[1] (n = 0.69) as compared to poly[2] (n = 0.08). In XPS analysis of both films, Li+ was also detected (from Li+ClO4− ion pairs) at very low levels, 0.26% or nondetected for poly[1] in the reduced and oxidized states, respectively, whereas for poly[2] the Li% is 1.55 and 0.31 for the reduced and oxidized states, respectively; this suggests that the oxidation charge compensation in this polymer film is to a certain degree attained by Li+ egress. However, due to the low sensitivity of XPS instrumentation for Li element, the small Li % have high inaccuracy, and consequently were not considered in the calculation of the values presented in Tables 1 and S3. Furthermore, it should be mentioned that the XPS results refer to typically 4−5 nm from the total film thickness. In Figure 3 are depicted SEM micrographs and EDS spectra for poly[1] films on reduced and oxidized states; SEM images of the surface of poly[2] films (not shown) reveal a similar morphology. The films consisted of a continuous layer (Z2), on top of which there are irregular fragments of higher dimensions (Z1). The EDS spectra showed that, in the Z1 regions, the Ni element (from polymeric film) is present in higher quantity than In (from substrate), while, in Z2 regions, the In content increases considerably in comparison to Ni. These results indicate that the film surface and thickness are not uniform. The EDS spectra revealed yet a higher Cl content in oxidized film (in agreement with XPS results), and the respective

Figure 1. CVs of the electrodeposition of (a) [Ni(3-Mesalen)] and (b) [Ni(3-MesaltMe)] complexes in ITO/PET, using 1.0 mmol dm−3 solutions of complexes in LiClO4/CH3CN 0.1 mol dm−3, at 0.02 V s−1 during 10 cycles. Inset: Chemical structures of the respective [Ni(salen)] complex.

in LiClO4/PC electrolytes. In contrast, electrochemical responses of poly[2] obtained in LiClO4/PC present greater peak separation (anodic peaks at more positive potentials and the cathodic peaks at less positive potentials) than those obtained in LiClO4/CH3CN. The different redox response of poly[2] in the two electrolyte solutions can be related to the monomer ligand structure; in the case of poly[2] the imine bridge presents four methyl groups, which may have different conformations in the two used electrolyte solutions, leading to distinct supramolecular structures, that may induce different electrochemical responses. We note additional anodic charge in the response to the first voltammetric cycle in background electrolyte. This is attributed to (irreversible) completion of the polymerization process for modest amounts of monomer and/or oligomer trapped within the film. In general, the voltammetric responses of polymeric films stabilized after 5 scans, which correspond to the typical “film conditioning”.43,45 The coulometric data revealed Γ = 0.31 and 1.13 μmol cm−2 for poly[1] and poly[2], respectively, suggesting greater polymerization efficiency for poly[2]. 3.2. Compositional and Morphological Characterizations. Surface analysis of poly[Ni(salen)] films in reduced and oxidized states (at E = 0.0 V and E = 1.3 V) was performed 14234

DOI: 10.1021/acsami.6b01977 ACS Appl. Mater. Interfaces 2016, 8, 14231−14243

Research Article

ACS Applied Materials & Interfaces

Figure 2. Voltammetric responses of (a) poly[1] (Γ = 0.31 μmol cm−2) and (b) poly[2] (Γ = 1.13 μmol cm−2) films in LiClO4/CH3CN (panel A) and LiClO4/PC (panel B) 0.1 mol dm−3 electrolytes, acquiring at the scan rate of 0.02 V s−1.

Table 1. XPS Results for Poly[1]: Surface Atomic Percentages and Atomic Ratios for Reduced and Oxidized Polymeric Films in LiClO4/CH3CN and LiClO4/PC 0.1 mol dm−3 atomic % films

Ni

C

atomic ratios

N

reduced state oxidized state

3.31 1.76

78.60 60.52

7.89 6.06

reduced state oxidized state

2.87 2.51

76.41 71.31

6.93 6.55

O LiClO4/CH3CN 10.03 26.33 LiClO4/PC 12.66 17.39

micrograph showed a bulkier film, possibly due to the ingress of supporting electrolyte into the polymeric matrix. 3.3. UV−Vis Spectroscopy in Situ. In Figure 4a and b are depicted the absolute UV−vis spectra acquired during the oxidations of poly[1] films in LiClO4/CH3CN and LiClO4/PC 0.1 mol dm−3, respectively; the equivalent spectra for poly[2] are depicted in Figure S2a and b; the spectra obtained during films reduction showed an inverse behavior and were omitted for simplicity. The poly[1] spectra show four bands in both electrolytes. For poly[2], four electronic bands are observed in LiClO4/ CH3CN, whereas in LiClO4/PC five bands are observed. The bands show different dependences on applied potential, and so their intricate behaviors are most readily appreciated by depicting the difference spectra, using the responses at selected

Cl

N/Ni

O/Ni

Cl/Ni

0.16 5.32

2.38 3.44

3.03 14.96

0.05 3.02

0.87 2.23

2.41 2.61

4.41 6.93

0.30 0.89

potentials as reference points, Figures S3 and S4 for poly[1] and poly[2], respectively. The absorbance versus potential profiles (Abs vs E) for wavelengths of detected bands (Figure 4a′ and b′ and Figure S2a′ and b′) allowed the identification of three main band profiles with increasing potential in both electrolytes: (i) bands at λ = 327 and 330 nm for poly[1] and poly[2], respectively, showed a monotonically intensity decrease during the positive half cycle, (ii) bands in the range λ = 402−415 nm and λ = 843 and higher than 950 nm showed intensity increase until E = 0.8−0.9 V and thereafter decreased until the end of the positive half cycle, and (iii) bands in the range λ = 510−525 nm for poly [1] and poly[2] and at λ = 639 nm (poly[2] in LiClO4/PC) showed monotonically intensity increase from E = 0.8−0.9 V until the end of the positive half scan. Curiously, the maximum intensity of the electronic bands 14235

DOI: 10.1021/acsami.6b01977 ACS Appl. Mater. Interfaces 2016, 8, 14231−14243

Research Article

ACS Applied Materials & Interfaces

Figure 3. SEM micrographs and respective EDS spectra at selected zones for poly[1] films in (a) reduced and (b) oxidized states, with a magnification of 20 000 times.

electronic bands at λ = 327/330 nm (3.79 and 3.76 eV) are assigned to the intervalence band (band gap), because they represent the largest energy transition and decrease in intensity upon polymer oxidation; (ii) the bands at λ = 402/415 nm (3.09 and 2.99 eV) are attributed to the transition between the valence band and the antibonding polaron level; and (iii) the bands at λ = 843 nm and λ > 950 nm (1.47 and 1.31 eV) are ascribed to the transition from the bonding to the antibonding polaron levels (or from the valence band to the bonding polaron level). The other band predicted by the polaronic model was estimated to be at λ = 766 and λ = 738 nm (1.62 and 1.68 eV) for poly[1] and poly[2], respectively; these would most likely be overlapped with the bands at λ = 843 and λ > 950 nm and thus not separately observed. Last, the very intense bands at λ = 510 nm (2.43 eV), λ = 520/525 nm (2.39/2.36 eV), and λ = 639 nm (1.94 eV) (for poly[2] in LiClO4/PC) showed different Abs versus E profiles and consequently cannot be assigned to the same charge carriers and explained by the polaronic model. Because the band absorbance starts to increase at high oxidation potentials and their ε-values depend on the solvent (CH3CN vs PC), unlike the other observed electronic transitions, they are assigned to charge transfer transitions between the metal and the oxidized ligand.32,40 3.4. Electrochromic Properties Evaluation. The visual inspection of both poly[Ni(salen)] films during redox cycling allowed the observation of their color changes. Figure 5 shows

at λ = 510 nm for poly[1] and at λ = 520/525 nm for poly[2] is significantly lower in LiClO4/PC than in LiClO4/CH3CN. Table 2 summarizes data for the observed electronic bands for both polymers in the two electrolytes: λmax (nm), molar extinction coefficients, ε (estimated from eq 1, through the slopes of Abs versus Q plots, Figure S5), and E (eV). Both polymers showed similar energy values for all of the electronic bands as a consequence of their similar monomer structures; furthermore, the electronic band energies for each polymer film are also similar in the two electrolytes, indicating that the polymer structures are comparable in both electrolytes. The results also revealed that, for each film, all of the electronic bands have similar ε-values in LiClO4/CH3CN and LiClO4/PC electrolytes, with the exception of the band at λ = 510 nm for poly[1] and λ = 520/525 nm for poly[2], the εvalues of which are lower in LiClO4/PC than in LiClO4/ CH3CN (for poly[1], the ε-value decreased from 19.7 × 103 to 13.6 × 103 cm−1 mol−1 dm3). For both polymers, all of the electronic band ε-values are typical of electronic transitions between states with large contribution from ligand orbitals, which is consistent with the assertion of ligand-based film oxidation.32,40,46 Consequently, the electronic bands can be assigned according to the polaronic model.32,40 The electronic bands with the same Abs versus E (or Q) profiles can be associated with the same charge carriers, whereby the following band assignment is proposed: (i) the 14236

DOI: 10.1021/acsami.6b01977 ACS Appl. Mater. Interfaces 2016, 8, 14231−14243

Research Article

ACS Applied Materials & Interfaces

Figure 4. Panel A: Absolute UV−visible spectra of poly[1] (Γ = 0.16 μmol cm−2) acquired during film oxidation in 0.1 mol dm−3 (a) LiClO4/ CH3CN and (b) LiClO4/PC (referenced to respective electrolytes spectra; orange - spectra at −0.1 V, green- spectra at 1.3 V). Panel B: Abs versus E plots for electronic bands identified in absolute UV−vis spectra, referenced to spectra at −0.1 V (arrows indicate scan direction).

Table 2. Electronic Bands and Respective Molar Extinction Coefficients (ε) for Poly[1] and Poly[2] in LiClO4/CH3CN and LiClO4/PC film

electrolyte solution

poly[1]

LiClO4/CH3CN

LiClO4/PC

poly[2]

LiClO4/CH3CN

LiClO4/PC

λmax/nm (eV) 327 402 510 843 327 402 510 843 330 415 525 >950 330 415 520 639 >950

(3.79) (3.09) (2.43) (1.47) (3.79) (3.09) (2.43) (1.47) (3.76) (2.99) (2.36) (1.31) (3.76) (2.99) (2.39) (1.94) (1.31)

ε × 10−3/mol−1 dm3 cm−1 4.99 3.06 19.66 5.21 5.90 3.12 13.64 6.72 0.76 0.42 4.22 0.68 0.60 0.31 1.37 0.64 0.54

Figure 5. Photographs of (a) poly[1] and (b) poly[2] films in different oxidation states (0.0, 0.7, and 1.3 V vs Ag/AgCl (NaCl/1.0 mol dm−3)) in LiClO4/CH3CN electrolyte.

photographic images of the film distinct colors as a function of the potential: in the reduced state (0.0 V) the films are yellow, switch to green at 0.7 V, and then to russet (reddish-brown) at 1.3 V. To explore the potential application of both films as EC materials, a square-wave potential step method coupled with UV−vis spectroscopy was used to evaluate their EC properties: switching times, optical contrasts, changes of the optical

densities, and coloration efficiencies. In Figure 6 are depicted the chronoabsorptograms obtained for poly[1] films (Γ = 0.16 μmol cm−2) in LiClO4/CH3CN and LiClO4/PC, for the two color changes, yellow ↔ green (λ = 750 nm) and green ↔ russet (λ = 510 nm). The results obtained for poly[2] (Γ = 0.84 μmol cm−2) in a similar study are depicted in Figure S6, and in 14237

DOI: 10.1021/acsami.6b01977 ACS Appl. Mater. Interfaces 2016, 8, 14231−14243

Research Article

ACS Applied Materials & Interfaces

time values obtained in LiClO4/CH3CN are τ = 9/10 s for the yellow ↔ green transition for both poly[1] and poly[2] and τ = 10/13 and 8/9 s for the green ↔ russet transition of poly[1] and poly[2], respectively. In LiClO4/PC, the switching time values are τ = 13/14 s for poly[1] and τ = 14/15 s for poly[2], in the case of the yellow ↔ green transition, whereas τ = 14/24 and 9/11 s for the green ↔ russet transition of poly[1] and poly[2], respectively, being clearly higher than those determined in LiClO4/CH3CN. The higher switching times observed in LiClO4/PC can be related to the lower degree of PC film solvation that may hamper the ClO4− counterions ingress, necessary to the charge balance during redox switching. This assumption is in agreement with the XPS results (Table 1), which revealed a larger increase in O/Ni ratio, from the reduced to the oxidized state, in LiClO4/CH3CN than in LiClO4/PC. Furthermore, independent of the electrolyte solution employed, the color transition yellow ↔ green is the fastest for poly[1], while for poly[2] is the green ↔ russet transition. The switching time values suggested that these films are especially useful for applications in which the demand for rapid switching is smaller, as in slow-acting electronic devices,12 such as smart cards or labels. As a complementary result, in Figure S8 are depicted, as an example, the chronoamperograms/chronocoulograms obtained simultaneously with the chronoabsorptograms for the color transition yellow ↔ green of poly[1] film in LiClO4/PC. It is possible to see that the switching time values estimated from current intensity and charge are similar to the values determined from absorbance. In fact, from the results, it is seen that the absorbance variation with the redox switching is reflected in the chronoamperograms by a current decay as the potential become constant (indicating that at each time a portion of the film is undergoing a redox process);19 the charge represents the same information in integrated form. The strong couple between these parameters highlights the possibility of the use of the current intensity/charge profiles as accurate parameters to evaluate the switching times of these polymeric films. In Table 3 are summarized the optical contrasts, changes of the optical density, charges requirements, and coloration efficiencies values, determined for the two films color changes in both electrolyte solutions. The optical contrast, defined as the transmittance difference, ΔT%, between redox states,23 was measured from the chronoabsorptograms, considering the full optical change. For poly[1] film in LiClO4/CH3CN, the optical contrast for the color transition yellow ↔ green is higher than that for the green ↔ russet transition (ΔT = 44.9% vs 32.7%). In LiClO4/PC, the optical contrasts decrease for both color

Figure 6. Chronoabsorptograms for poly[1] films in LiClO4/CH3CN (black) and LiClO4/PC (blue), for the color transitions (a) yellow ↔ green (λ = 750 nm) and (b) green ↔ russet (λ = 510 nm), with indication of the estimated switching times.

Figure S7 are depicted the UV−vis spectra for the two films in each color and in both supporting electrolytes, with identification of the considered wavelengths. The switching times (τ), estimated from absorbance−time curves and indicated in Figures 6 and S6, were determined considering 90% of the full optical change.19,21 The switching

Table 3. EC Parameters for Poly[1] and Poly[2] Films in LiClO4/CH3CN and LiClO4/PC: Optical Contrasts (ΔT%), Changes of the Optical Density (ΔOD), Charge Requirements (Qd), and Coloration Efficiencies (η) film

electrolyte solution

color transitiona

ΔT%

ΔOD

Qd/mC cm−2

η/cm2 C−1

poly[1]

LiClO4/CH3CN

yellow ↔ green green ↔ russet yellow ↔ green green ↔ russet yellow ↔ green green ↔ russet yellow ↔ green green ↔ russet

44.9 32.7 26.2 18.3 13.0 19.4 51.2 58.5

0.47 0.92 0.49 0.53 0.11 0.36 0.13 0.32

5.62 6.05 6.46 4.92 2.99 2.42 4.02 3.80

83.18 152.04 75.55 107.36 36.94 146.86 33.14 84.12

LiClO4/PC poly[2]

LiClO4/CH3CN LiClO4/PC

a

λ = 510 nm and λ = 750 nm for the color transitions green ↔ russet and yellow ↔ green, respectively. 14238

DOI: 10.1021/acsami.6b01977 ACS Appl. Mater. Interfaces 2016, 8, 14231−14243

Research Article

ACS Applied Materials & Interfaces changes (ΔT = 26.2% and 18.3%, respectively). The poly[2] film has an opposite behavior: the color transition green ↔ russet has a higher contrast than the yellow ↔ green transition (ΔT = 19.4% vs 13.0%), being that in LiClO4/PC the contrasts are enhanced (for ΔT = 58.5% and 51.2%). The measured optical contrasts are in accordance with the color difference observed in photographs of Figure 5. The change of the optical density, ΔOD, was estimated using eq 2. For both polymeric films, the ΔOD values for the yellow ↔ green color transition are very similar in LiClO4/CH3CN and LiClO4/PC. For the green ↔ russet transitions, on the other hand, the ΔOD are significantly lower in LiClO4/PC (for poly[1], ΔOD = 0.92 in LiClO4/CH3CN and 0.53 in LiClO4/ PC). These results reflect the behavior of the obtained UV−vis spectra represented in Figures 4, S2, or S7. The smaller ΔOD value in LiClO4/PC is in accordance with the decrease of the intensity of the band at λ = 510−525 from LiClO4/CH3CN to the LiClO4/PC electrolyte solution (as discussed in section 3.3), providing an indication that the green ↔ russet color change is mainly associated with the appearance of the band at λ = 510−525 nm, attributed to CT transitions. For the remaining UV−vis bands, no significant differences were observed in the two electrolyte solutions, which is consistent with the similar ΔOD values for the yellow ↔ green color change. The power efficiency of the EC films was measured through the coloration efficiency, η, determined by eq 3, considering the Qd values obtained from the chronocoulograms, as represented in Figure S8b for poly[1] in LiClO4/PC. The η values measured are in the range 33.14−83.18 cm2 C−1 for the color transition yellow ↔ green and 84.12−152.04 for the green ↔ russet transition. The higher values for the green ↔ russet transition can be explained by the induced higher ΔOD, with similar or even lower charge requirements. Furthermore, for the same color change, the films presented smaller η values in LiClO4/PC electrolyte. Comparing both films, poly[1] showed better coloration efficiency. However, it is important to note that eq 3 is typically used for the characterization of bleached to colored switching materials. Consequently, EC materials that exhibit transitions within the visible region, as in these films, tend to lower the η values, although they can reflect into modest optical changes.12 The electrochemical stability of poly[Ni(salen)] films was evaluated by chronoamperometric observations, investigating the influence of the electrolyte solutions, LiClO4/CH3CN or LiClO4/PC, on the cycle life of yellow ↔ green and green ↔ russet color transitions. The obtained chronoamperograms are depicted in Figure 7 for poly[1] and in Figure S9 for poly[2]; in the figures are also indicated the percentage of charge loss, corresponding to the number of redox cycles after which no further color change was observable. Independent of the color change, both polymeric films showed greater electrochemical stability in LiClO4/PC than in LiClO4/CH3CN. The best stability was obtained for poly[1] film in LiClO4/PC that changed color between yellow and green during around 9000 redox cycles with a charge loss of 34.3%. The same film, in LiClO4/CH3CN, showed color change only during around 2000 redox cycles, with a charge loss of 60.0%. The green ↔ russet color transition was less stable: after 400 redox cycles, poly[1] showed a loss in charge of 96.1% in LiClO4/CH3CN and 68.2% in LiClO4/PC. Moreover, for the yellow ↔ green change, poly[2] showed charge losses of 10.5% and 70.9% in LiClO4/PC and LiClO4/

Figure 7. Chronoamperograms of poly[1] films in LiClO4/CH3CN (black) and LiClO4/PC (blue) for the color changes (a) yellow ↔ green and (b) green ↔ russet, applying two potential pulses of 50 s by redox cycle, with potential alternating between (a) 0.0−0.7 V and (b) 0.7−1.3 V.

CH3CN, respectively, at around 650 redox cycles (taking in account that, for this film, the color contrast between the yellow and green is low). The green ↔ russet change, although showing the highest contrast for this film, was also revealed to be unstable, with charge losses of 70.2% in LiClO4/PC and 88.6% in LiClO4/CH3CN (at 650 cycles). These results showed that poly[2] has an inferior electrochemical stability performance as compared to poly[1]. Furthermore, the reasons for the general better stabilities in LiClO4/PC are not wellknown, but can be related to the less volatility of PC as compared to CH3CN and to degradation of the substrate (PET) over time, when immersed in CH3CN. In Table S4 are summarized the switching times determined for both poly[Ni(salen)] films during the stability studies, along the redox cycles. Note that these response times are in general higher than those indicated on Figures 6, S6, and S8, because the films used here were deposited during 10 scans (Γ ≈ 0.31 14239

DOI: 10.1021/acsami.6b01977 ACS Appl. Mater. Interfaces 2016, 8, 14231−14243

Research Article

ACS Applied Materials & Interfaces

Figure 8. Pictures (panel A) and schematic illustrations (panel B) of the assembled ECDs of (a and a′) typology 1 and (b and b′) typology 2, showing the color contrast between the two electrodes, each one in different oxidation states: electrode 1 in reduced state (−1.0 and −1.1 V) and electrode 2 in oxidized state (1.0 and −0.25 V); in (b′) the cut to prevent the electrical contact corresponds to the butterfly’s shape/borders, and the connection to the potential source is made through the link between the butterfly’s wing and the clean surface of ITO.

and 1.13 μmol cm−2, for poly[1] and poly[2], respectively), while the films used in chronoabsorptometric studies were deposited during 5 scans (Γ ≈ 0.16 and 0.84 μmol cm−2, for poly[1] and poly[2], respectively) and so have different thicknesses. In a general way, the response times for the color changes yellow−green and green−russet (oxidation processes) tend to increase slightly or remain nearly constant, while the response times for the color changes green−yellow and russet−green (reduction processes) tend to decrease slightly or remain constant. The electrochemical stability of poly[1] film in LiClO4/PC was also evaluated by CV during 200 redox cycles (Figure S10). The voltammetric responses showed that the anodic peak corresponding to the process associated with the color change from the yellow to green (Epa = 0.37 V) shifted to more positive potentials along the successive redox cycles, suggesting morphological and/or structural changes in film matrix. On the other hand, the anodic peak corresponding to the process associated with the color change from green to russet (Epa = 0.99 V) showed a more rapid decrease in current intensity, reaching very low values. At this point, the anodic peak at more positive potential values was overlapped by the peak associated with the yellow to green color change, such that this became the dominant observed color change. These observations can justify the lower stability of the green ↔ russet color change. Taking into account the diversity of ligands already used in the design of poly[Ni(salen)] films, it can be concluded that the high electrochemical stability obtained for poly[1] results from the presence of σ-donor methyl groups in the 3-position of the salicylaldehyde moieties; furthermore, the presence of four methyl groups in the imine bridge had a negative effect on the electrochemical stability of previously studied poly[Ni(salen)] films. It can be recognized that the electrochemical stability of poly[Ni(salen)] films resulted from a subtle combination of

substituent effects in the imine bridge and/or in the 3-position of the salicylaldehyde moiety: substituents in the 5-position of salicylaldehyde moieties prevent polymerization, and positions 4- and 6- have a lower effect on the overall stability because they are not directly conjugated with the C−O−Ni moiety. Although no chronoamperometric studies have been made for previously studied poly[Ni(salen)] films, the voltammetric studies suggested that poly[Ni(salen)] (pristine salen, no substituents) showed lower stability than poly[Ni(saltMe)]47 (pristine salen with four methyl groups in the imine bridge, σdonors) and poly[Ni(3-MeOsaltMe)]48 (pristine saltMe with methoxy group, π donor and σ-withdraw group, in the 3position of the salicylaldehyde moiety). In this context, the choice of salen ligands with a σ-donor methyl group in the 3position of the salicylaldehyde moieties, individually, and combined with four methyl groups in the imine bridge could account for high stability in the oxidized state, provided by the electron σ-donor properties of the methyl groups. However, the four methyl groups in the imine bridge may also have a significant impact on the supramolecular polymer structure (difficult to anticipate) due to the different conformations for the ethylene backbone in the imine bridge, which can favor/ disfavor the film electrochemical stability; in the present study, it disfavored poly[2] electrochemical stability. As a summary, the high electrochemical stability obtained for poly[1] in LiClO4/PC electrolyte solution constitutes a step forward in the electrochemical stability improvement when compared to other poly[M(salen)] films,35,36 and highlights the potentiality of this film as a very stable EC material. Furthermore, when compared to other EC materials, poly[1] in LiClO4/PC showed faster responses (lower switching times) than the inorganic metal oxides (typically, τ = 12−60 s),37,42 and similar or even higher optical contrasts and coloration efficiencies than some conducting polymers and derived composites.20,38,49 14240

DOI: 10.1021/acsami.6b01977 ACS Appl. Mater. Interfaces 2016, 8, 14231−14243

Research Article

ACS Applied Materials & Interfaces 3.5. ECDs Fabrication and Characterization. The high electrochemical stability of the color change yellow ↔ green of the poly[1], allied with its good optical contrast, makes it the choice for the fabrication of a solid-state ECD; in the case of poly[2], although the high optical contrast associated with the green ↔ russet color change could be a key issue in the selection for ECD fabrication, its low electrochemical stability remains the highest important drawback. Thereby, ECDs were fabricated using poly[1] as EC material using a lateral configuration, in two figures of merit, a simple shape (typology 1) and a butterfly shape (typology 2), the latter to exploit the possibility of a future application in smart cards or labels. In Figure 8 are presented the pictures and schematic illustrations of the assembled ECDs. In both typologies, the electrodes are in the same plane and are designated by 1 or 2: the electrode 1 is in the reduced state and showed a yellow color, while electrode 2 is in the oxidized state and exhibited a green color. In Figure 9a are depicted the chronoamperograms obtained during the study of the electrochemical stability of the ECD of typology 1, considering the color change yellow ↔ green. The ECD retained a clear color change for ca. 3000 redox cycles, with a loss in charge of 37.0%. During the first 1000 redox cycles, the process was simultaneously monitored by in situ UV−vis spectroscopy, at λ = 750 nm (Figure 9b). At the end of these cycles, the ECD registered a charge loss of 23.1% and a loss of 15.8% in absorbance. The estimated τ values are reported in Figure 9c. From the absorbance−time curves, τ = 157 s for the color change yellow−green and τ = 145 s for the reverse color change. These times are similar to those determined from the chronoamperograms: τ = 151 and 142 s for the yellow−green and green−yellow color changes, respectively. Furthermore, the optical contrast of the assembled ECD was calculated as ΔT = 88.7% (at λ = 750 nm) and the coloration efficiency as η = 130.4 cm2 C−1, with ΔOD = 0.84 and a charge requirement of Qd = 6.4 mC cm−2. To evaluate the application of this EC system in devices with more sophisticated template forms, a proof-of-concept ECD of typology 2 was assembled. The electrochemical stability of the ECD of typology 2 was determined through the chronoamperograms depicted in Figure S11. The ECD changed color between yellow and green during around 1250 redox cycles, with a charge loss of 58.9%. These results indicate the need for a more optimized typology when more sophisticated ECD templates are needed in terms of future applications.

4. CONCLUSIONS Poly[1] and poly[2] films were successfully electrodeposited on ITO/PET flexible substrates. Voltammetric characterization revealed that poly[1] showed similar redox profiles in LiClO4/ CH3CN and LiClO4/PC, while poly[2] showed solventdependent electrochemical responses. Both films showed multielectrochromic behavior, exhibiting yellow, green, and russet colors, according to their oxidation state. EC properties were first studied in solution with both polymer films showing switching times in the range 8−24 s; this is consistent with their potential application in slow-acting electronic devices. For poly[1], the yellow ↔ green color transition showed the best optical contrast, whereas for poly[2] it was the green ↔ russet transition. Poly[1] also showed the best coloration efficiencies, with the highest values being obtained for the green ↔ russet color change. Both polymeric

Figure 9. ECD of typology 1: (a) chronoamperograms obtained by the application of two potential pulses of 200 s by redox cycle, with potential alternating between −1.0 V (yellow) and 1.0 V (green); chronoabsorptograms (at λ = 750 nm) and respective chronoamperograms obtained during (b) the first 1000 redox cycles and (c) the first three redox cycles, with indication of the switching times.

films were more electrochemically stable in LiClO4/PC, although the overall poorer electrochemically stability of 14241

DOI: 10.1021/acsami.6b01977 ACS Appl. Mater. Interfaces 2016, 8, 14231−14243

Research Article

ACS Applied Materials & Interfaces

Stability of Their Electrochromic Devices (PPy/PANI-PEDOT, PANI/PPy-PEDOT). Electrochim. Acta 2013, 96, 214−224. (8) Saikia, D.; Pan, Y. C.; Wu, C. G.; Fang, J.; Tsai, L. D.; Kao, H. M. Synthesis and Characterization of a Highly Conductive OrganicInorganic Hybrid Polymer Electrolyte Based on Amine Terminated Triblock Polyethers and its Application in Electrochromic Devices. J. Mater. Chem. C 2014, 2 (2), 331−343. (9) Sydam, R.; Deepa, M.; Shivaprasad, S. M.; Srivastava, A. K. A WO3-Poly(Butyl Viologen) Layer-by-Layer Film/Ruthenium Purple Film Based Electrochromic Device Switching by 1 V Application. Sol. Energy Mater. Sol. Cells 2015, 132, 148−161. (10) Osterholm, A. M.; Shen, D. E.; Kerszulis, J. A.; Bulloch, R. H.; Kuepfer, M.; Dyer, A. L.; Reynolds, J. R. Four Shades of Brown: Tuning of Electrochromic Polymer Blends Toward High-Contrast Eyewear. ACS Appl. Mater. Interfaces 2015, 7 (3), 1413−1421. (11) Li, K. R.; Zhang, Q. H.; Wang, H. Z.; Li, Y. G. Red, Green, Blue (RGB) Electrochromic Fibers for the New Smart Color Change Fabrics. ACS Appl. Mater. Interfaces 2014, 6 (15), 13043−13050. (12) Beaujuge, P. M.; Reynolds, J. R. Color Control in π-Conjugated Organic Polymers for Use in Electrochromic Devices. Chem. Rev. 2010, 110 (1), 268−320. (13) Argun, A. A.; Reynolds, J. R. Line Patterning for Flexible and Laterally Configured Electrochromic Devices. J. Mater. Chem. 2005, 15 (18), 1793−1800. (14) Thakur, V. K.; Ding, G. Q.; Ma, J.; Lee, P. S.; Lu, X. H. Hybrid Materials and Polymer Electrolytes for Electrochromic Device Applications. Adv. Mater. 2012, 24 (30), 4071−4096. (15) Schott, M.; Lorrmann, H.; Szczerba, W.; Beck, M.; Kurth, D. G. State-of-the-Art Electrochromic Materials Based on Metallo-Supramolecular Polymers. Sol. Energy Mater. Sol. Cells 2014, 126, 68−73. (16) Kraft, A.; Rottmann, M. Properties, Performance and Current Status of the Laminated Electrochromic Glass of Gesimat. Sol. Energy Mater. Sol. Cells 2009, 93 (12), 2088−2092. (17) Gillaspie, D. T.; Tenent, R. C.; Dillon, A. C. Metal-Oxide Films for Electrochromic Applications: Present Technology and Future Directions. J. Mater. Chem. 2010, 20 (43), 9585−9592. (18) Sydam, R.; Deepa, M.; Joshi, A. G. A Novel 1,1 ′-Bis 4-(5,6Dimethyl-1H-Benzimidazole-1-yl)Butyl-4,4’-Bipyridinium Dibromide (Viologen) for a High Contrast Electrochromic Device. Org. Electron. 2013, 14 (4), 1027−1036. (19) Amb, C. M.; Dyer, A. L.; Reynolds, J. R. Navigating the Color Palette of Solution-Processable Electrochromic Polymers. Chem. Mater. 2011, 23 (3), 397−415. (20) Yu, W. Y.; Chen, J.; Fu, Y. L.; Xu, J. K.; Nie, G. M. Electrochromic Property of a Copolymer Based on 5-Cyanoindole and 3,4-Ethylenedioxythiophene and its Application in Electrochromic Devices. J. Electroanal. Chem. 2013, 700, 17−23. (21) Wu, T. Y.; Liao, J. W.; Chen, C. Y. Electrochemical Synthesis, Characterization and Electrochromic Properties of Indan and 1,3Benzodioxole-Based Poly(2,5-Dithienylpyrrole) Derivatives. Electrochim. Acta 2014, 150, 245−262. (22) Rende, E.; Kilic, C. E.; Udum, Y. A.; Toffoli, D.; Toppare, L. Electrochromic Properties of Multicolored Novel Polymer Synthesized Via Combination of Benzotriazole and N-Functionalized 2,5-di(2Thienyl)-1H-Pyrrole Units. Electrochim. Acta 2014, 138, 454−463. (23) Chang, K. H.; Wang, H. P.; Wu, T. Y.; Sun, I. W. Optical and Electrochromic Characterizations of Four 2,5-Dithienylpyrrole-Based Conducting Polymer Films. Electrochim. Acta 2014, 119, 225−235. (24) Jarosz, T.; Brzeczek, A.; Walczak, K.; Lapkowski, M.; Domagala, W. Multielectrochromism of Redox States of Thin Electropolymerised Films of Poly(3-Dodecylpyrrole) Involving a Black Coloured State. Electrochim. Acta 2014, 137, 595−601. (25) Kurtay, G.; Ak, M.; Gullu, M.; Toppare, L.; Ak, M. S. Synthesis and Electropolymerization of 3,4-Substituted Quinoxaline Functionalized Pyrrole Monomer and Optoelectronic Properties of its Polymer. Synth. Met. 2014, 194, 19−28. (26) Lu, J. L.; Liu, W. S.; Ling, H.; Kong, J. H.; Ding, G. Q.; Zhou, D.; Lu, X. H. Layer-by-layer Assembled Sulfonated-Graphene/ Polyaniline Nanocomposite Films: Enhanced Electrical and Ionic

poly[2]: the highest electrochemical stability was obtained for the yellow ↔ green change of poly[1] in LiClO4/PC, with charge loss of only 34.3% after 9000 cycles. Consequently, poly[1] was used to fabricate ECD with two figures of merit (typology 1 and 2), with typology 1 showing the best performance: ΔT = 88.7% (at λ = 750 nm), η = 130.4 cm2 C−1, and charge intensity decrease of 37.0% after 3000 redox cycles. The results showed that poly[1] has promising EC properties both in a solution cell and in a solid-state device. The combination of high electrochemical stability, interesting poly electrochromism, and good optical contrast motivates the application of poly[1] as EC material in electronic displays. Furthermore, the combination of these EC properties with a flexible substrate has considerable advantages that may be exploited in paper-like applications.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsami.6b01977. Summary of the peak potentials observed in CVs, XPS data for poly[2] and XPS spectra in O 1s region, UV−vis data for poly[2], differential UV−vis spectra and Abs versus Q plots, chronoabsorptograms/amperograms/ coulograms for poly[1] and poly[2], estimated switching times, and chronoamperograms of the ECD of typology 2 (PDF)



AUTHOR INFORMATION

Corresponding Author

*Tel.: +351 22 04020590. Fax: +351 22 0402 695. E-mail: [email protected]. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS We thank Fundaçaõ para a Ciência e a Tecnologia (FCT, Portugal) and FEDER under Programme PT2020 (UID/QUI/ 50006/2013). M.N. (SFRH/BD/79171/2011) and M.A. (SFRH/BD/89156/2012) also thank FCT for their grants.



REFERENCES

(1) Deb, S. K. A Novel Electrophotographic System. Appl. Opt. 1969, 8 (Suppl 1), 192−195. (2) Lian, W. R.; Huang, Y. C.; Liao, Y. A.; Wang, K. L.; Li, L. J.; Su, C. Y.; Liaw, D. J.; Lee, K. R.; Lai, J. Y. Flexible Electrochromic Devices Based on Optoelectronically Active Polynorbornene Layer and Ultratransparent Graphene Electrodes. Macromolecules 2011, 44 (24), 9550−9555. (3) Caglar, A.; Cengiz, U.; Yildirim, M.; Kaya, I. Effect of Deposition Charges on the Wettability Performance of Electrochromic Polymers. Appl. Surf. Sci. 2015, 331, 262−270. (4) Reyes-Gil, K. R.; Stephens, Z. D.; Stavila, V.; Robinson, D. B. Composite WO3/TiO2 Nanostructures for High Electrochromic Activity. ACS Appl. Mater. Interfaces 2015, 7 (4), 2202−2213. (5) Granqvist, C. G. Electrochromics for Smart Windows: OxideBased Thin Films and Devices. Thin Solid Films 2014, 564, 1−38. (6) Mortimer, R. J.; Varley, T. S. In Situ Spectroelectrochemistry and Colour Measurement of a Complementary Electrochromic Device Based on Surface-Confined Prussian Blue and Aqueous Solution-Phase Methyl Viologen. Sol. Energy Mater. Sol. Cells 2012, 99, 213−220. (7) Abaci, U.; Guney, H. Y.; Kadiroglu, U. Morphological and Electrochemical Properties of PPy, PAni Bilayer Films and Enhanced 14242

DOI: 10.1021/acsami.6b01977 ACS Appl. Mater. Interfaces 2016, 8, 14231−14243

Research Article

ACS Applied Materials & Interfaces Conductivities and Electrochromic Properties. RSC Adv. 2012, 2 (28), 10537−10543. (27) Sydam, R.; Kokal, R. K.; Deepa, M. Fast, Direct, Low-Cost Route to Scalable, Conductive, and Multipurpose Poly(3,4-Ethylenedixoythiophene)-Coated Plastic Electrodes. ChemPhysChem 2015, 16 (5), 1042−1051. (28) Xiong, S. X.; Fu, J. L.; Li, Z. F.; Shi, Y. J.; Wang, X. Q.; Chu, J.; Gong, M.; Wu, B. H. Modulating the Electrochromic Performances of Transmissive and Reflective Devices Using N, N-Dimethyl Formamide Modified Poly(3,4-Ethylenedioxythiophene)/Poly(Styrene Sulfonate) Blend as Active Layers. J. Macromol. Sci., Part B: Phys. 2015, 54 (7), 799−810. (29) Kang, J. H.; Oh, Y. J.; Paek, S. M.; Hwang, S. J.; Choy, J. H. Electrochromic Device of PEDOT-PANI Hybrid System for Fast Response and High Optical Contrast. Sol. Energy Mater. Sol. Cells 2009, 93 (12), 2040−2044. (30) Sydorov, D.; Duboriz, I.; Pud, A. Poly(3-Methylthiophene)Polyaniline Couple Spectroelectrochemistry Revisited for the Complementary Red-Green-Blue Electrochromic Device. Electrochim. Acta 2013, 106, 114−120. (31) Kraft, A.; Rottmann, M.; Gilsing, H. D.; Faltz, H. Electrodeposition and Electrochromic Properties of N-Ethyl Substituted Poly (3,4-Ethylenedioxypyrrole). Electrochim. Acta 2007, 52 (19), 5856− 5862. (32) Fonseca, J.; Tedim, J.; Biernacki, K.; Magalhaes, A. L.; Gurman, S. J.; Freire, C.; Hillman, A. R. Structural and Electrochemical Characterisation of Pd(salen)-Type Conducting Polymer Films. Electrochim. Acta 2010, 55 (26), 7726−7736. (33) Vilas-Boas, M.; Santos, I. C.; Henderson, M. J.; Freire, C.; Hillman, A. R.; Vieil, E. Electrochemical Behavior of a New Precursor for the Design of Poly Ni(salen)-Based Modified Electrodes. Langmuir 2003, 19 (18), 7460−7468. (34) Martins, M.; Boas, M. V.; de Castro, B.; Hillman, A. R.; Freire, C. Spectroelectrochemical Characterisation of Copper Salen-Based Polymer-Modified Electrodes. Electrochim. Acta 2005, 51 (2), 304− 314. (35) Pinheiro, C.; Parola, A. J.; Pina, F.; Fonseca, J.; Freire, C. Electrocolorimetry of Electrochromic Materials on Flexible ITO Electrodes. Sol. Energy Mater. Sol. Cells 2008, 92 (8), 980−985. (36) Branco, A.; Pinheiro, C.; Fonseca, J.; Tedim, J.; Carneiro, A.; Parola, A. J.; Freire, C.; Pina, F. Solid-State Electrochromic Cells Based on M(salen)-Derived Electroactive Polymer Films. Electrochem. SolidState Lett. 2010, 13 (9), J114−J118. (37) Kalagi, S. S.; Mali, S. S.; Dalavi, D. S.; Inamdar, A. I.; Im, H.; Patil, P. S. Limitations of Dual and Complementary Inorganic-Organic Electrochromic Device for Smart Window Application and its Colorimetric Analysis. Synth. Met. 2011, 161 (11−12), 1105−1112. (38) Xu, G. Q.; Zhao, J. S.; Cui, C. S.; Hou, Y. F.; Kong, Y. Novel Multicolored Electrochromic Polymers Containing Phenanthrene9,10-Quinone and Thiophene Derivatives Moieties. Electrochim. Acta 2013, 112, 95−103. (39) Freire, C.; de Castro, B. Spectroscopic Characterisation of Electrogenerated Nickel(III) Species. Complexes with N2O2 SchiffBase Ligands Derived from Salicylaldehyde. J. Chem. Soc., Dalton Trans. 1998, No. 9, 1491−1497. (40) Tedim, J.; Patricio, S.; Fonseca, J.; Magalhaes, A. L.; Moura, C.; Hillman, A. R.; Freire, C. Modulating Spectroelectrochemical Properties of Ni(salen) Polymeric Films at Molecular Level. Synth. Met. 2011, 161 (9−10), 680−691. (41) Dale, S. M.; Glidle, A.; Hillman, A. R. Spectroelectrochemical Observation of Poly(Benzo-Thiophene) N-Doping and P-Doping. J. Mater. Chem. 1992, 2 (1), 99−104. (42) Sapp, S. A.; Sotzing, G. A.; Reynolds, J. R. High Contrast Ratio and Fast-Switching Dual Polymer Electrochromic Devices. Chem. Mater. 1998, 10 (8), 2101−2108. (43) Tedim, J.; Carneiro, A.; Bessada, R.; Patricio, S.; Magalhaes, A. L.; Freire, C.; Gurman, S. J.; Hillman, A. R. Correlating Structure and Ion Recognition Properties of Ni(salen)-Based Polymer Films. J. Electroanal. Chem. 2007, 610 (1), 46−56.

(44) Tedim, J.; Goncalves, F.; Pereira, M. F. R.; Figueiredo, J. L.; Moura, C.; Freire, C.; Hillman, A. R. Preparation and Characterization of Poly Ni(salen) (Crown Receptor)/Multi-Walled Carbon Nanotube Composite Films. Electrochim. Acta 2008, 53 (23), 6722−6731. (45) Dahm, C. E.; Peters, D. G.; Simonet, J. Electrochemical and Spectroscopic Characterization of Anodically Formed Nickel Salen Polymer Films on Glassy Carbon, Platinum, and Optically Transparent Tin Oxide Electrodes in Acetonitrile Containing Tetramethylammonium Tetrafluoroborate. J. Electroanal. Chem. 1996, 410 (2), 163−171. (46) Vilas-Boas, M.; Freire, C.; de Castro, B.; Christensen, P. A.; Hillman, A. R. Spectroelectrochemical Characterisation of Poly Ni(saltMe)-Modified Electrodes. Chem. - Eur. J. 2001, 7 (1), 139−150. (47) Vilas-Boas, M.; Freire, C.; de Castro, B.; Hillman, A. R. Electrochemical Characterization of a Novel Salen-Type Modified Electrode. J. Phys. Chem. B 1998, 102 (43), 8533−8540. (48) Vilas-Boas, M.; Santos, I. C.; Henderson, M. J.; Freire, C.; Hillman, A. R.; Vieil, E. Electrochemical Behavior of a New Precursor for the Design of Poly[Ni(salen)]-Based Modified Electrodes. Langmuir 2003, 19 (18), 7460−7468. (49) Camurlu, P.; Gultekin, C.; Bicil, Z. Fast switching, High Contrast Multichromic Polymers from Alkyl-Derivatized Dithienylpyrrole and 3,4-Ethylenedioxythiophene. Electrochim. Acta 2012, 61, 50−56.

14243

DOI: 10.1021/acsami.6b01977 ACS Appl. Mater. Interfaces 2016, 8, 14231−14243